PDF File - Colorado State University

Mar 8, 2002 - Christoph Kiefl,† Narasimha Sreerama,† Raid Haddad,‡ Lisong Sun,‡ ... heme.1 In sperm-whale myoglobin, isomer A is the predomina...
0 downloads 0 Views 651KB Size
Published on Web 03/08/2002

Heme Distortions in Sperm-Whale Carbonmonoxy Myoglobin: Correlations between Rotational Strengths and Heme Distortions in MD-Generated Structures Christoph Kiefl,† Narasimha Sreerama,† Raid Haddad,‡ Lisong Sun,‡ Walter Jentzen,‡ Yi Lu,‡ Yan Qiu,‡ John A. Shelnutt,*,‡ and Robert W. Woody*,† Contribution from the Department of Biochemistry and Molecular Biology, Colorado State UniVersity, Fort Collins, Colorado 80523, and Biomolecular Materials and Interfaces Department, Sandia National Laboratories, Albuquerque, New Mexico 87185-1349, and Department of Chemistry, The UniVersity of New Mexico, Albuquerque, New Mexico 87131 Received August 13, 2001. Revised Manuscript Received December 26, 2001

Abstract: We have investigated the effects of heme rotational isomerism in sperm-whale carbonmonoxymyoglobin using computational techniques. Several molecular dynamics simulations have been performed for the two rotational isomers A and B, which are related by a 180° rotation around the R-γ axis of the heme, of sperm-whale carbonmonoxy myoglobin in water. Both neutron diffraction and NMR structures were used as starting structures. In the absence of an experimental structure, the structure of isomer B was generated by rotating the heme in the structure of isomer A. Distortions of the heme from planarity were characterized by normal coordinate structural decomposition and by the angle of twist of the pyrrole rings from the heme plane. The heme distortions of the neutron diffraction structure were conserved in the MD trajectories, but in the NMR-based trajectories, where the heme distortions are less well defined, they differ from the original heme deformations. The protein matrix induced similar distortions on the hemes in orientations A and B. Our results suggest that the binding site prefers a particular macrocycle conformation, and a 180° rotation of the heme does not significantly alter the protein’s preference for this conformation. The intrinsic rotational strengths of the two Soret transitions, separated according to their polarization in the heme plane, show strong correlations with the ruffling deformation and the average twist angle of the pyrrole rings. The total rotational strength, which includes contributions from the chromophores in the protein, shows a weaker correlation with heme distortions.

Introduction

Heme b (protoheme IX) in hemoproteins can exist in two different orientations1 that differ by a 180° rotation of the heme group about the R-γ axis (Figure 1), and the two rotational isomers are called isomers A and B. For isomer A, the porphyrin substituents as conventionally numbered (Figure 1) are arranged in a clockwise sense when viewed from the side of the proximal His for globins and other hemoproteins with a single protein ligand, or the N-terminal ligand for bis-liganded hemoproteins (e.g., cytochrome b5). For isomer B, the porphyrin substituents are arranged in a counterclockwise sense. Twofold rotation about the R-γ axis interchanges the 2- and 4-vinyl groups with the 3- and 1-methyl groups, respectively, and leads to different protein-heme contacts. NMR studies have shown that the equilibrium ratio of A to B form depends on the protein sequence, and on the oxidation state and the ligands of the heme.1 In sperm-whale myoglobin, isomer A is the predominant form. It has been shown that heme distortions can modulate * Corresponding authors. Robert W. Woody: Phone:(970) 491-6214; Fax (970) 491-0494, e-mail: [email protected]. John A. Shelnutt: Phone (505) 272-7160; Fax (505) 272-7077, e-mail: [email protected]. † Colorado State University ‡ Sandia National Laboratories and University of New Mexico. (1) La Mar, G. N.; Toi, H.; Krishnamoorthi, R. J. Am. Chem. Soc. 1984, 106, 6395-6401. 10.1021/ja011961w CCC: $22.00 © 2002 American Chemical Society

Figure 1. Isomers A and B, which differ by a 180° rotation of the heme about the R-γ axis in the protein matrix. This isomerization exchanges the position of the 2-vinyl with the 3-methyl and the 4-vinyl with the 1-methyl group. Vinyl groups are depicted in the syn-orientation. The pyrrole rings are identified by letters A to D, and the methine carbons by Greek letters R to δ. The view is from the proximal side of the heme.

physiologically important properties of hemoproteins.2 Heme rotational isomerism could influence these physiologically important properties. Since the protein matrix can induce different distortions in isomers A and B, the heme in isomer B may exhibit different functional properties than that in isomer A. Experimental results initially suggested that isomer B of native sperm-whale myoglobin has a 10-fold higher affinity for O2 than isomer A.3 However, subsequent experiments did not J. AM. CHEM. SOC. 2002, 124, 3385-3394

9

3385

ARTICLES

Kiefl et al.

show any significant differences in the O2- and CO-binding of native and reconstituted sperm-whale or yellowfin-tuna myoglobin.4-6 The differences found in the first experiment were probably due to unrecognized irreversible oxidation of a fraction of heme, which resulted in higher apparent binding constants for O2.4 The O2 affinity is probably modulated mainly by steric and electronic effects depending on the peripheral substituents.7,8 However, the Bohr effect in Chironomus thummi thummi hemoglobins is affected by the heme rotational equilibrium.8 The major form (isomer A) shows a large Bohr effect, whereas the minor form (isomer B) exhibits little or no Bohr effect. The heme isomers of bovine cytochrome b5 have also been shown to differ in a physiologically relevant parameter: isomer A has a reduction potential that is 27 mV more negative than that of isomer B.9 Native sperm-whale myoglobin exhibits heme isomerism, with ∼10% of isomer B at equilibrium. Many high-resolution crystal structures have been reported for myoglobin, but only the A isomer has been observed in crystal structures10 and NMR structures.11 Freshly reconstituted myoglobin consists of a 50: 50 mixture of the two isomers and reaches equilibrium at a rate that depends on several factors, such as species, pH, temperature, etc.1 The Soret CD spectrum of freshly reconstituted myoglobin is only half as intense as that of the native protein, suggesting that isomer B has a weak, or even negative, Soret CD.5,12 Extrapolation of the linear relation between the Soret CD intensity and the isomeric composition of the reconstituted carbonmonoxy myoglobin, as measured by NMR, gave a ∆max of +90 and -7 M-1 cm-1 respectively for isomers A and B.13 To examine the origins of differences in the Soret CD of the two isomers of myoglobin, we have performed MD simulations of sperm-whale myoglobin in both A and B forms. The structure of the B isomer, which is not available experimentally, was generated from the structure of isomer A by rotating the heme about the R-γ axis in the heme pocket. The calculated Soret CD spectra of the two isomers are qualitatively similar. The origins of the Soret CD lie in the distortions of the heme, leading to an intrinsic rotational strength, and the interaction between the heme chromophore and the chromophores of the protein matrix.14 In this paper we present the analysis of the protein-induced distortions of the heme in the MD trajectories for isomers A and B using normal-coordinate structural decomposition15,16 (NSD) and another simple measure of heme distortion. We correlate the heme deformations with the (2) Shelnutt, J. A.; Song, X. Z.; Ma, J. G.; Jia, S. L.; Jentzen, W.; Medforth, C. J. Chem. Soc. ReV. 1998, 27, 31-41. (3) Livingston, D. J.; Davis, N. L.; La Mar, G. N.; Brown, W. D. J. Am. Chem. Soc. 1984, 106, 3025-3026. (4) Aojula, H. S.; Wilson, M. T.; Morrison, I. E. G. Biochem. J. 1987, 243, 205-210. (5) Light, W. R.; Rohlfs, R. J.; Palmer, G.; Olson, J. S. J. Biol. Chem. 1987, 262, 46-52. (6) Neya, K.; Funasaki, N.; Shiro, Y.; Iizuka, T.; Imai, K. Biochim. Biophys. Acta 1994, 1208, 31-37. (7) Chang, C. K.; Ward, B.; Ebina, S. Arch. Biochem. Biophys. 1984, 231, 366-371. (8) Gersonde, K.; Sick, H.; Overkamp, M.; Smith, K. M.; Parish, D. W. Eur. J. Biochem. 1986, 157, 393-404. (9) Walker, F. A.; Emrick, D.; Rivera, J. E.; Hanquet, B. J.; Buttlaire, D. H. J. Am. Chem. Soc. 1988, 110, 6234-6240. (10) Cheng, X.; Schoenborn, B. P. J. Mol. Biol. 1991, 220, 381-399. (11) O ¨ sapay, K.; Theriault, Y.; Wright, P. E.; Case, D. J. Mol. Biol. 1994, 244, 183-197. (12) Aojula, H. S.; Wilson, M. T.; Drake, A. Biochem. J. 1986, 237, 613-616. (13) Aojula, H. S.; Wilson, M. T.; Moore, G. R.; Williamson, D. J. Biochem. J. 1988, 250, 853-858. (14) Hsu, M. C.; Woody, R. W. J. Am. Chem. Soc. 1971, 93, 3515-3525. 3386 J. AM. CHEM. SOC.

9

VOL. 124, NO. 13, 2002

calculated rotational strengths. The ruffling deformation16 and the average twist angle of the opposite pyrrole rings, which also measures the ruffling deformation, show a strong correlation with the intrinsic rotational strengths of the two Soret transitions of the heme. The total rotational strength, which includes coupling with other chromophores in the protein, shows a weak correlation. A preliminary account of this work was presented at an Alcuin Symposium.17 Methods MD-Simulation. Molecular dynamics simulations of the two rotational isomers of sperm-whale carbonmonoxy myoglobin in water have been performed at 300 K with GROMOS96.18 The starting structures of the protein were taken from a neutron-diffraction structure10 (PDB entry 2mb5) and two of twelve NMR structures11 (PDB entry 1myf). After placing the protein in a rectangular box of equilibrated extended simple-point-charge (SPC/E) water19 molecules and removing water molecules within a 2.3 Å radius of any non-hydrogen protein atoms, two cycles of steepest-descent energy minimization were performed. In the first cycle, the positions of the protein atoms were restrained, but in the second, no restraints were applied. All covalent bond lengths and the H2O bond angle in this and subsequent steps were constrained to their equilibrium values with SHAKE,20 with a geometric tolerance of 10-4. Eight water molecules were replaced by Cl- ions to maintain electroneutrality, followed by two cycles of steepest-descent energy minimization, as described above. This was followed by 20 steps of conjugate-gradient energy minimization. The molecular dynamics simulation under periodic boundary conditions was then initiated, starting at 100 K, and using an integration time step of 2 fs. Three steps followed, each of 5 ps duration, in which the temperature was increased by 100 K per step, and restraints were decreased. Three more steps, each of 5 ps duration, followed at 100, 200, and 300 K, without restraints. Finally, unrestrained simulation at 300 K for 20 ps completed the equilibration process. Nonbonded interactions were calculated with a triple-range method,21 with cutoffs of 8 and 14 Å. Coordinates were saved at 0.050-ps intervals. Further details are provided in the Supporting Information. We performed four MD simulations for each isomer, with four different initial velocity assignments, using the neutron-diffraction structure10 as the starting geometry. The starting geometry for isomer B was obtained from the isomer A structure by rotation of the heme by 180° about the R-γ axis, keeping the vinyl groups in plane and in the syn conformation, as shown in Figure 1. The four MD trajectories of isomer A, generated with the neutron-diffraction structure and different initial velocities, are referred to as T1A, T2A, T3A, and T4A. The corresponding trajectories of isomer B are referred to as T1BT4B. We have also performed MD simulations with two starting structures taken from the ensemble of twelve NMR structures,11 structures #1 (1myf-1) and #8 (1myf-8), and these MD trajectories are referred to as N1 and N2, respectively, with the isomer indicated by A and B, e.g., N1A for isomer A with NMR structure #1 as the starting geometry. These MD simulations were performed on two of the original (15) Jentzen, W.; Ma, J. G.; Shelnutt, J. A. Biophys. J. 1998, 74, 753-763. (16) Jentzen, W.; Song, X. Z.; Shelnutt, J. A. J. Phys. Chem. B 1997, 101, 16841699. (17) Woody, R. W.; Kiefl, C.; Sreerama, N.; Lu, Y.; Qiu, Y.; Shelnutt, J. A. In Insulin and Related Proteins: From Structure to Function and Pharmacology; Federwisch, M., Dieken, M. L., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2002 (in press). (18) van Gunsteren, W. F.; Billeter, S. R.; Eising, A. A.; Hu¨nenberger, P. H.; Kru¨ger, P.; Mark, A. E.; Scott, W. R. R.; Tironi, I. G. Biomolecular Simulations: The GROMOS96 Manual and User Guide; Hochschulverlag AG an der ETH: Zu¨rich, 1996. (19) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. J. Phys. Chem. 1987, 91, 6269-6271. (20) Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. J. Comput. Phys. 1977, 23, 327-341. (21) van Gunsteren, W. F.; Berendsen, H. J. C. Angew. Chem., Int. Ed. Engl. 1990, 29, 992-1023.

Heme Distortions in SW-Myoglobin

ARTICLES

NMR structures instead of using the average of the twelve NMR structures. In this way, we covered more conformational space and additionally could monitor the dependence of MD-generated structures on initial distortions. The trajectories N1 and N2 have different starting geometries, whereas trajectories T1 to T4 have the same starting geometry but differ in their initial velocities. For comparison with the myoglobin MD simulations, a 600-ps MD simulation of Fe(III) protoporphyrin chloride (FePP(Cl)) was performed by using the POLYGRAPH 3.21 program and a DREIDING II force field (MSI) modified to include atom types for the heme with fractional atomic charges as previously described.2,16 The MD calculation was run by using canonical NVT dynamics at 300 K while saving snapshot structures at 0.05-ps intervals. The snapshots from the FePP(Cl) MD trajectory were analyzed in terms of the normal coordinate deformations in the same way as the myoglobin MD trajectories. Intrinsic Rotational Strength. Circular dichroism (∆) is the differential absorption of left and right circularly polarized light,

∆ ) L - R

(1)

where L and R are respectively the molar decadic extinction coefficients for left- and right-circularly polarized light. The rotational strength of a transition from the ground state 0 to excited state a is determined by integrating the intensity under a single band in the CD spectrum, according to:

R0a ) (3000pc ln10/16π2NA)



band

(∆/λ) dλ ) 0.2477



band

(∆/λ) dλ (2)

where p is Planck’s constant divided by 2π, c is the velocity of light, and NA is Avogadro’s number. Rotational strength is usually expressed in Debye-Bohr magnetons (1 DBM ) 0.9273 × 10-38 cgs units). The numerical factor in the second form of eq 2 provides R in DBM when ∆ is in M-1 cm-1. The rotational strength can be calculated theoretically22 from the imaginary part of the scalar product of the electric and magnetic dipole transition moments:

R0a ) Im{(0|µ|a)‚(a|m|0)}

(3)

where µ is the electric dipole transition moment operator, a measure of the linear displacement of charge upon excitation; and m is the magnetic dipole transition moment operator, a measure of the circular displacement of electron density upon excitation. The superposition of µ and m results in a helical displacement of charge, which interacts differently with left- and right-circularly polarized light. There are two types of contributions to the rotational strengths of the two Soret components: the intrinsic rotational strength of the transitions resulting from the inherent chirality of the heme, and the perturbations resulting from coupling of the heme transitions with transitions in other chromophores in the protein, of which the aromatic side chains and the backbone peptide groups are the most important. In this paper, we consider only the intrinsic rotational strengths of the Soret components. The extrinsic contributions arising from coupling of these transitions to proteins groups are treated in another paper (Kiefl et al., in preparation). The intrinsic rotational strengths for the Soret components were calculated according to the equation:

R0a ) -(e2p2/4πm2cν0a)(0|∇|a)‚(a|r × ∇|0)

(4)

where (0|∇|a) is the dipole velocity matrix element connecting the ground state, 0, and the excited state, a; (a|r × ∇|0) is a matrix element connecting the ground state, 0, and the excited state, a, that is (22) Rosenfeld, L. Z. Phys. 1928, 52, 161-174.

proportional to the angular momentum of the transition 0 f a; ν0a is the frequency (s-1) of the transition 0 f a; and e and m are respectively the charge and mass of the electron. This formulation of the rotational strength, known as the dipole velocity form, is origin-independent,23 in contrast to the dipole length form in eq 3. In these calculations we treated the heme as a porphyrin dianion, considering only the π electrons, neglecting the methyl and propionyl side chains, but including the two vinyl groups. The π-system in the heme group contained 28 atoms. The transition parameters and intrinsic rotational strength of the Soret transitions were calculated by using π-molecular orbital theory in the Pariser-Parr-Pople (PPP) approximation24 with parameters from the “standard model” of Weiss et al.25 The calculations were performed on the heme structures in the MD trajectories at 0.05-ps intervals. In this paper, we will only consider the intrinsic rotational strengths of the heme. The total rotational strengths of the heme transitions, including contributions of coupling with transitions in protein chromophores, are described elsewhere (Kiefl et al., in preparation). The Soret CD band has two nearly degenerate components that are polarized in the plane of the heme and perpendicular to each other. The near-degeneracy of the Soret components leads to fluctuations in the relative energies of the x- and y-polarized Soret components, precluding a unique correlation using the energy classification. The two Soret components were distinguished according to their polarization in the heme-plane, which gives a good correlation with a measure of heme distortion as described below. Each Soret transition moment was assigned as either x- or y-polarized, according to the angle of its projection on the mean plane of the heme, using (45° lines as boundaries. The x-axis passes through the nitrogen atoms of the pyrroles D and B (ND- - -NB) and the y-axis through A and C pyrroles (NA- - NC) (Figure 1). The calculated rotational strengths for the x- and y-polarized components were separately averaged over the trajectory. Normal-Coordinate Structural Analysis. Heme structures were analyzed by normal-coordinate structural decomposition (NSD) as described earlier.15,16 Figure 2 illustrates the NSD method for the heme from the X-ray crystal structure of horse heart cytochrome c. The computational procedure projects out the deformations along all of the 66 normal coordinates of the porphyrin macrocycle, including both inplane and out-of-plane modes. Typically for hemoprotein crystal and NMR structures, however, the deformations are only large enough to be statistically significant for the lowest frequency out-of-plane modes. This is because these modes have the smallest restoring forces and thus tend to have the largest deformations from planarity. For porphyrin crystal structures, which are usually refined to much higher resolution than hemoproteins, the lowest frequency modes alone do not provide a sufficient description of the structure,15 but for the hemoproteins, a description in terms of the deformations of only the lowest frequency normal modes is adequate statistically. That is, the error in the structure when the contributions from all of the other modes are neglected is less than the positional uncertainty in the X-ray structure. For this work, a C version of the NSD program was modified for input of the 12 000 snapshot structures from the MD simulation. Timeaveraged plots were obtained by using the 2D smoothing function of SigmaPlot2001, which uses a Gaussian weighting function over a 6-ps sliding window and a linear fitting function. The fast Fourier transform of the POWSPEC program of SigmaPlot was applied to the deformations in the first 8192 (213) structures of the trajectories T2A, T2B, and FePP(Cl). Deformations along each of the normal modes were weighted by a cosine function that vanishes at the endpoints of the data (Hanning window26), thus removing high-frequency components generated by using a finite trajectory. The magnitudes resulting from (23) Moffitt, W. J. Chem. Phys. 1956, 25, 467-478. (24) (a) Pariser, R.; Parr, R. G. J. Chem. Phys. 1953, 21, 466-471; 767-776. (b) Pople, J. A. Trans. Faraday Soc. 1953, 49, 1375-1385. (25) Weiss, C.; Kobayashi, H.; Gouterman, M. J. Mol. Spectrosc. 1965, 16, 415-450. J. AM. CHEM. SOC.

9

VOL. 124, NO. 13, 2002 3387

ARTICLES

Kiefl et al. deformation also gives rise to opposite signs of τAC and τBD, but the average propellering deformation is much smaller than the ruffling deformation.

Results and Discussion

Figure 2. Illustration of the heme distortions characterized by normalcoordinate structural decomposition of the heme from horse heart cytochrome c. The distorted structure at the top left is that of the heme in horseheart cytochrome c. Each of the six structures below it shows the heme deformed along one of the lowest-frequency normal modes. Each deformation belongs to an irreducible representation of the D4h symmetry group appropriate to the planar, 4-fold symmetric porphyrin nucleus. A linear combination of these deformations, called saddling (sad), ruffling (ruf), doming (dom), waving-x (waVx), waving-y (waVy), and propellering (pro), can be fit to an arbitrarily deformed heme within the experimental error of protein crystal structures. The coefficients in the linear combination required to fit horse-heart cytochrome c are represented in the bar plot on the right. These coefficients are the displacements in Å along these six normal coordinates that best represent the out-of-plane distortion of the heme structure.

the transform were squared and multiplied by two to yield the power spectral density as a function of frequency. Twist-Angle (τ). The virtual dihedral angles between the two sets of opposing pyrrole rings can be used to characterize the overall distortion of the heme. These relate the pyrrole rings ATC and BTD (see Figure 1) and are denoted τAC and τBD, respectively. The dihedral angle τAC is determined by sighting along the line connecting the midpoints of pyrroles A and C. The angle carrying the C2B-C3B bond of pyrrole A into the C4B-C5B bond of pyrrole C is τAC. An analogous procedure is used to define τBD. A planar heme has both τAC and τBD equal to 0°. These two dihedral angles give a measure of the twist of the pyrrole rings with respect to the heme plane and are inversely related. Generally, a positive AC twist (0° < τAC < 90°) is compensated by a negative BD twist (-90° < τAC < 0°), which is a consequence of the constraints of the heme macrocycle. For small values of τAC and τBD, the inverse relation may not be satisfied, but the heme deformations are also small in such cases. The average twist angle (τ) was calculated by averaging the absolute values of the two dihedral angles and retaining the sign of the AC twist angle. The average twist angle shows a strong correlation with the ruffling deformation from the NSD analysis. The propellering (26) Press, W. H.; Flannery, B. P.; Teukolsky, S. A.; Vetterling, W. T. Numerical Recipes: The Art of Scientific Computing; Cambridge University Press: Cambridge UK, 1989; p 425. 3388 J. AM. CHEM. SOC.

9

VOL. 124, NO. 13, 2002

Heme Distortions: Normal Coordinate Structural Analysis. The major heme deformations that make up the distortion in hemoproteins are ruffling (ruf), saddling (sad), and doming (dom), whereas waVing (waV(x), waV(y)) and propellering (pro) deformations tend to be smaller due to their higher vibrational frequencies15,16 (Figure 2). These six macrocycle deformations are unique because they correspond to the lowest frequency outof-plane vibrational modes of B1u, B2u, A2u, Eg, and A1u symmetries (irreducible representations), respectively, of the nominal D4h molecular symmetry of hemes. Using NSD, the out-of-plane distortion of most heme structures can be well represented in terms of the normal coordinate displacements for these six vibrational modes. The out-of-plane heme distortions in myoglobin are generally small in comparison with other hemoproteins.15 For the 14 experimental structures of sperm-whale carbonmonoxy-myoglobin obtained by X-ray diffraction (1vxf), neutron diffraction (2mb5), and NMR spectroscopy (1myf: 12 structures), the total out-of-plane distortions are generally less than 0.7 Å. The consistent features of the experimental heme structures from all sources are the small, positive doming, y-waving, and propellering deformations (